Hostname: page-component-848d4c4894-nr4z6 Total loading time: 0 Render date: 2024-06-07T23:25:39.498Z Has data issue: false hasContentIssue false

Open-system behaviour of detrital zircon during weathering: an example from the Palaeoproterozoic Pretoria Group, South Africa

Published online by Cambridge University Press:  14 December 2021

Tom Andersen*
Affiliation:
Department of Geology, University of Johannesburg, PO Box 524, Auckland Park, 2006, Johannesburg, South Africa Department of Geosciences, University of Oslo, PO Box 1047 Blindern, N-0316Oslo, Norway
Marlina A. Elburg
Affiliation:
Department of Geology, University of Johannesburg, PO Box 524, Auckland Park, 2006, Johannesburg, South Africa
*
Author for correspondence: Tom Andersen, Email: tom.andersen@geo.uio.no
Rights & Permissions [Opens in a new window]

Abstract

Detrital zircon in six surface samples of sandstone and contact metamorphic quartzite of the Magaliesberg and Rayton formations of the Pretoria Group (depositional age c. 2.20–2.06 Ga) show a major age fraction at 2.35–2.20 Ga, and minor early Palaeoproterozoic – Neoarchaean fractions. Trace-element concentrations vary widely, with Ti, Y and light rare earth elements (LREEs) spanning over three orders of magnitude. REE distribution patterns range from typical zircon patterns (LREE depletion, heavy REE enrichment, well-developed positive Ce and negative Eu anomalies) to patterns that are flat to concave downwards, with indistinct Ce and Eu anomalies. The change in REE pattern correlates with increases in alteration-sensitive parameters such as Ti concentration and (Dy/Sm) + (Dy/Nd), U–Pb discordance and content of common lead, and with a gradual washing-out of oscillatory zoning in cathodoluminescence images. U and Th concentrations also increase, but Th/U behaves erratically. Discordant zircon scatters along lead-loss lines to zero-age lower intercepts, suggesting that the isotopic and chemical variations are the results of disturbance long after deposition. The rocks sampled have been in a surface-near position (at least) since Late Cretaceous time, and exposed to deep weathering under intermittently hot and humid conditions. In this environment, even elements commonly considered as relatively insoluble could be mobilized locally, and taken up by radiation-damaged zircon. Such secondary alteration effects on U–Pb and trace elements can be expected in zircon in any ancient sedimentary rock that has been exposed to tropical–subtropical weathering, which needs to be considered when interpreting detrital zircon data.

Type
Original Article
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NC
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial licence (http://creativecommons.org/licenses/by-nc/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the original article is properly cited. The written permission of Cambridge University Press must be obtained prior to any commercial use.
Copyright
© The Author(s), 2021. Published by Cambridge University Press

1. Introduction

Crystalline zircon is a robust and non-reactive mineral that can survive abrasion during repeated events of erosion and transport, and whose U–Pb system can be preserved even at high-grade metamorphic conditions (e.g. Williams, Reference Williams2001; Bindeman et al. Reference Bindeman, Schmitt, Lundstrom and Hervig2018). The crystal structure of zircon will, however, suffer radiation damage from the decay of U and Th incorporated at the time of crystallization, and their radioactive decay products. Over geological time, this will cause gradual transformation of the mineral into an amorphous substance known as metamict zircon. Whereas radiation damage itself does not cause changes in chemical composition or discordance of the U–Pb isotope system, metamict zircon is mechanically weakened (e.g. Salje, Reference Salje2006) and reactive when exposed to fluids, for example during diagenesis or weathering (Balan et al. Reference Balan, Neuville, Trocellier, Fritsch, Muller and Calas2001; Willner et al. Reference Willner, Sindern, Metzger, Ermolaeva, Kramm, Puchkov and Kronz2003; Hay & Dempster, Reference Hay and Dempster2009; Pidgeon et al. Reference Pidgeon, Nemchin and Cliff2013, Reference Pidgeon, Nemchin, Roberts, Whitehouse and Belluci2019; Andersen et al. Reference Andersen, Elburg and Van Niekerk2019 b). When interacting with fluids, metamict zircon commonly loses radiogenic lead causing normal U–Pb discordance; in the process, the contents of common lead and non-structural elements such as titanium, light rare earth elements (REEs) and hydrogen will increase (Stern et al. Reference Stern, Goldich and Newell1966; Black, Reference Black1987; Nasdala et al. Reference Nasdala, Wenzel, Vavra, Irmer, Wenzel and Kober2001; Belousova et al. Reference Belousova, Griffin, O’Reilly and Fisher2002; Hoskin & Schaltegger, Reference Hoskin and Schaltegger2003). Even elements such as yttrium, uranium and thorium can be introduced during weathering of metamict zircon (Pidgeon et al. Reference Pidgeon, Nemchin and Whitehouse2017, Reference Pidgeon, Nemchin, Roberts, Whitehouse and Belluci2019).

The favourable properties of crystalline zircon have made U–Pb ages from detrital zircon in clastic sediments a much-used tool for provenance identification and correlation of sedimentary strata (e.g. Zimmermann, Reference Zimmermann, Siegesmund, Basei, Oyhantcabal and Oriolo2018). Furthermore, the trace-element signature can be used to identify the types of igneous rocks contributing material to a basin (Belousova et al. Reference Belousova, Griffin, O’Reilly and Fisher2002; Griffin et al. Reference Griffin, Belousova, Shee, Pearson and O’Reilly2004; Veevers & Saeed, Reference Veevers and Saeed2007). Zircon grains whose composition has been modified during diagenesis or weathering of the host sediment are less useful for these purposes. When dating igneous or metamorphic rocks by U–Pb in zircon, grains that are influenced by secondary processes can commonly be avoided by careful selection of single grains for analysis, and altered parts of grains can be removed by mechanical or chemical abrasion (Krogh, Reference Krogh1982; Mattinson, Reference Mattinson2005). In detrital zircon geochronology, the priority is to establish an unbiased estimate of the zircon population in the samples studied. This is commonly achieved by random sampling from a bulk zircon mineral separate, and selective methods used in isotope dilution–thermal ionization mass spectrometry (ID-TIMS) U–Pb geochronology are generally not applicable. For meaningful interpretations of sedimentary provenance, stratigraphic correlation or basin filling history to be extracted from detrital zircon data, it is important that no significant age fraction is overlooked and no spurious fraction added to the dataset. To preserve such ‘qualitative representativity’ (in the sense of Andersen et al. Reference Andersen, Elburg and Magwaza2019 a), it is essential that the observed age, isotopic or trace-element distributions are not modified by alteration processes after deposition, that is, that detrital zircon has behaved as a closed system. Different data filtering criteria have been suggested to exclude analyses that have been compromised by alteration processes from detrital zircon datasets, the most common of which is to remove analyses that deviate from the U–Pb concordia curve by more than a given percentage (a discordance filter). This has the disadvantage that it may also exclude grains which are discordant only due to recent lead loss, whose 207Pb/206Pb age still retains a valid memory of the age of the protosource rock, whereas it may be inefficient against bias-inducing lead loss caused by processes in the past (Andersen et al. Reference Andersen, Elburg and Magwaza2019 a). Other filtering criteria that have been proposed are based on trace-element parameters (Bell et al. Reference Bell, Boehnke, Barboni and Harrison2019) or common lead content (Andersen et al. Reference Andersen, Elburg and Van Niekerk2019 b). The main concern when using such data filters should be to remove grains whose U–Pb isotope systematics have been affected, while retaining as much of the valid information in the randomly sampled dataset as possible.

The vulnerability of zircon to alteration depends on the degree of structural damage, which is related to the alpha radiation dose accumulated over the lifetime of a zircon grain. At time t, a crystal formed at t i  > t will have accumulated an alpha radiation dose given by:

(1) \begin{align*} {D_\alpha }(t) & = {{{N_A}} \over {{{10}^6}}}\bigg[{{8{A_{238}}{C_U}} \over {{M_{238}}}}({e^{{\lambda _{238}}{t_i}}} - {e^{{\lambda _{238}}t}}) + {{7{A_{235}}{C_U}} \over {{M_{235}}}}({e^{{\lambda _{235}}{t_i}}} - {e^{{\lambda _{235}}t}}) \\ & + {{6{C_{Th}}} \over {{M_{232}}}}({e^{{\lambda _{232}}{t_i}}} - {e^{{\lambda _{232}}t}})\bigg] \end{align*}

where A 235, A 238 are the natural isotopic abundances of 235U and 238U; M 235, M 238 and M 232 are the atomic masses of 235U, 238U and 232Th; C Th and C U are the concentrations (in parts per million) of U and Th; and N A is the Avogadro constant. The formula was first defined for t = 0 by Holland & Gottfried (Reference Holland and Gottfried1955); here it is generalized as a function of t from the version given by Nasdala et al. (Reference Nasdala, Reiners, Garver, Kennedy, Stern, Balan and Wirth2004). The corresponding weight fraction of metamict material (f m) in a zircon grain is given by the empirical relationship:

(2) $${f_m} = 1 - {e^{ - {B_\alpha }{D_\alpha }}},$$

in which B α  = 2.7 × 10–19 g/α (Zhang & Salje, Reference Zhang and Salje2001).

X-ray and spectroscopic studies have shown that the change from a fully crystalline to an amorphous state in zircon is a continuous transition process in which the crystal passes through a stage with crystalline ‘islands’ in a continuous, amorphous matrix, which is reached at what is known as the percolation point (Salje et al. Reference Salje, Chronsch and Ewing1999). A critical alpha dose of D α  ≈ 3.5 × 1018 α/g has been shown to be necessary to transform a fully crystalline zircon to this intermediate state (Salje et al. Reference Salje, Chronsch and Ewing1999), at which 61 weight percent of a grain will consist of metamict zircon according to Equation (2). Such zircon grains have lost much of their mechanical strength and resistance against chemical alteration; zircon with D α > 3.5 × 1018 α/g has been shown to be depleted in sediments relative to their source rocks (Markwitz & Kirkland, Reference Markwitz and Kirkland2018). Further radiation damage will lead to complete metamictization at D α ≥ 8 × 1018 α/g (Ewing et al. Reference Ewing, Meldrum, Wang, Weber and Corrales2003).

As an example of the extent of radiation damage expected in ancient zircon, the distribution of D α at the present day (D α (0)) expected from a suite of 2.2 Ga zircon with U and Th concentrations following the distribution patterns observed for granitic zircon by Belousova et al. (Reference Belousova, Griffin, O’Reilly and Fisher2002) is illustrated in Figure 1. In this example, only c. 22% of the zircon grains would be expected to contain less than 61% metamict material at the present time (D α (0) < 3.5 × 1018 α/g), whereas c. 50% of the grains would be completely metamict with D α (0) ≥ 8 × 1018 α/g. A similar distribution based on data for zircon in a wider range of igneous rocks (gabbro to alkali granite) by Kirkland et al. (Reference Kirkland, Smithies, Taylor, Evans and McDonald2015) gives a similar total range of D α , but with a lower median value and hence lower accumulated alpha dose (23% with D α (0) ≥ 3.5 × 1018 α/g, and 6% with D α (0) ≥ 8 × 1018 α/g).

Fig. 1. Distribution of accumulated alpha dose at t = 0 Ma experienced by 2.2 Ga zircon with U and Th concentration distributions similar to those reported for zircon in granitic rocks by Belousova et al. (Reference Belousova, Griffin, O’Reilly and Fisher2002), represented by percentile points and a log-normal distribution compatible with these, and for a wider compositional range of igneous rocks by Kirkland et al. (Reference Kirkland, Smithies, Taylor, Evans and McDonald2015); log-normal distribution based on data from their supplementary table A1. D α(0) values at which zircon would have reached the percolation point (3.5 × 1018 α/g) and the complete metamictization limit (8.0 × 1018 α/g) are as given by Salje et al. (Reference Salje, Chronsch and Ewing1999).

A consequence of this example is that a significant fraction of zircon grains from a 2.2 Ga granitic source rock would be unlikely to survive erosion, transport and deposition to make it into a sedimentary basin at the present time. Another consequence is that detrital zircons in old sedimentary rocks that were fully crystalline at the time of deposition may have acquired sufficient structural damage while residing in their host sedimentary rock, making their U–Pb and trace-element systems vulnerable to reaction with fluids in a near-surface weathering environment. Such damaged detrital grains may survive physically, as long as the host sediment is not eroded and recycled. However, both their U–Pb isotope composition and trace-element distributions may be modified.

To get a better understanding of the effects of post-depositional, in situ weathering on detrital zircon in ancient sedimentary rocks, and its consequences for detrital zircon geochronology, we have undertaken a U–Pb and trace-element study of zircon in samples of quartz arenites of the Magaliesberg and Rayton formations of the Palaeoproterozoic Pretoria Group in part of the Transvaal Basin of South Africa (Fig. 2). These rocks have remained undisturbed since deposition some time before 2.06 Ga (Zeh et al. Reference Zeh, Ovtcharova, Wilson and Schaltegger2015, Reference Zeh, Wilson and Ovtcharova2016, Reference Zeh, Wilson and Gerdes2020), but have been in a surface-near position and exposed to weathering since Late Cretaceous time (e.g. Partridge & Maud, Reference Partridge and Maud1987; Partridge, Reference Partridge1998).

Fig. 2. (a) Generalized stratigraphic column of the Pretoria Group in the south-central part of the Transvaal Basin, South Africa, simplified from Eriksson et al. (Reference Eriksson, Altermann, Hartzer, Johnson, Anhaeusser and Thomas2006, fig. 9). The Rooiberg Group lavas have been dated to 2061 ± 2 Ma by a lead evaporation age on zircon (Walraven, Reference Walraven1997), and the intrusive rocks of the Bushveld complex (BVC) by ID-TIMS U–Pb on zircon to 2056 ± 0.3 Ma by Zeh et al. (Reference Zeh, Ovtcharova, Wilson and Schaltegger2015). Further geochronological evidence limiting time of deposition of Pretoria Group strata are: (1) Timeball Hill Formation, syn-sedimentary ashlayers, Rasmussen et al. (Reference Rasmussen, Bekker and Fletcher2013); (2) younger ashlayers in the Timeball Hill Formation also provide the currently most robust available maximum limit for the age of the Hekpoort lavas; and (3) minimum age of the Daspoort and lower part of the Silverton formations sandstone is given by an Ar–Ar age on a cross-cutting mafic–ultramafic dyke swarm (Wabo et al. Reference Wabo, Humbert, de Kock, Belyanin, Söderlund, Maré, Beukes, Srivastava, Ernst and Peng2019). Sample numbers shown in parentheses (Magaliesberg and Rayton formations) refer to localities shown by circles in (b). (b) Simplified geological map of the south-central part of the Transvaal Basin, after Council of Geoscience 1: 250 000 geological mapsheets Rustenburg, Pretoria, West Rand and East Rand. The 500°C isograd of the Bushveld contact aureole (i) and the outer limit of the aureole (ii) are from Cawthorn et al. (Reference Cawthorn, Eales, Walraven, Uken, Watkeys, Johnson, Anhaeusser and Thomas2006). Extents of preserved African and Post African 1 surfaces are from Partridge (Reference Partridge1998). V – Vredefort Dome, centre of the 2.02 Ga Vredefort meteorite impact; PS – axis of the Potchefstroom Syncline, from Brink et al. (Reference Brink, Waanders, Bisschoff and Gay2000).

2. Geological setting

The late Archaean – Palaeoproterozoic Transvaal Supergroup (Eriksson et al. Reference Eriksson, Altermann, Hartzer, Johnson, Anhaeusser and Thomas2006) comprises shales, sandstones, carbonate rocks and minor volcanic rocks, deposited on a basement of Archaean gneisses and supracrustal cover successions of the Kaapvaal Craton. It is preserved in three different basins, the Transvaal and Griqualand West basins of South Africa and the Kanye Basin of Botswana. Correlations between the successions in these basins as outlined by, for example, Eriksson et al. (Reference Eriksson, Altermann, Hartzer, Johnson, Anhaeusser and Thomas2006), have been questioned from more recent geochronological data (Mapeo et al. Reference Mapeo, Armstrong, Kampunzu, Modisi, Ramokate and Modie2006; Moore et al. Reference Moore, Polteau, Armstrong, Corfu and Tsikos2012; Gumsley et al. Reference Gumsley, Chamberlain, Bleeker, Söderlund, de Kock, Larsson and Bekker2017).

The Pretoria Group is the youngest part of the Transvaal Supergroup in the Transvaal Basin; a simplified, general stratigraphic column is given in Figure 2a. The Pretoria Group comprises two unconformity-bounded sequences, deposited in fault-controlled basins on the Kaapvaal Craton (Eriksson et al. Reference Eriksson, Altermann, Hartzer, Johnson, Anhaeusser and Thomas2006). Only the younger of the two sequences is of interest to the present study. In the Transvaal Basin, this succession started with terrestrial sedimentation (Boshoek Formation) and andesitic volcanism (Hekpoort Formation), followed by lacustrine deposition (Dwaalheuwel and Strubenkop formations), marine transgression (Daspoort and Silverton formations) and regression during deposition of the Magaliesberg Formation. The Magaliesberg Formation consists of sandstone with minor lenses of mudrock, overlying the marine shale of the Silverton Formation. It has variously been interpreted as shallow-marine, or as a succession of regression-related shore, tidal and braided delta deposits (Eriksson et al. Reference Eriksson, Altermann, Hartzer, Johnson, Anhaeusser and Thomas2006). The younger, ‘Post-Magaliesberg’ formations of the Pretoria Group are represented by the Rayton Formation in the area of this study (Fig. 2), and by several formations in the eastern part of the Transvaal Basin (Schreiber & Eriksson, Reference Schreiber and Eriksson1992). These strata comprise mainly sandstones and shales, with subordinate carbonates and volcanic rocks, deposited in non-marine, probably isolated sub-basins (Eriksson et al. Reference Eriksson, Altermann, Catuneanu, van der Merwe and Bumby2001). In the southern exposure area in Figure 2b, in the Potchefstroom syncline, the Magaliesberg formation occurs as discontinuous erosional remnants, and stratigraphically higher units are not preserved.

Deposition of the Pretoria Group in the Transvaal Basin must have started slightly before 2300 Ma (Hannah et al. Reference Hannah, Bekker, Stein, Markey and Holland2004; Rasmussen et al. Reference Rasmussen, Bekker and Fletcher2013), and terminated before eruption of the lavas of the overlying Dullstroom Formation and Rooiberg Group, the latter of which has been dated to 2061 ± 2 Ma (Walraven, Reference Walraven1997). Deposition was followed by emplacement of the Bushveld Complex at 2056.0 ± 0.3 Ma (Zeh et al. Reference Zeh, Wilson and Ovtcharova2016) and the Vredefort meteorite impact at 2023 ± 4 Ma (Kamo et al. Reference Kamo, Reimold, Krogh and Colliston1996). Datable units within the Pretoria Group are few, and most are only imprecisely dated (Fig. 2a), so the depositional chronology has mainly been constrained by detrital zircon data (Dorland, Reference Dorland2004; Schröder et al. Reference Schröder, Beukes and Armstrong2016; Zeh et al. Reference Zeh, Wilson and Ovtcharova2016, Reference Zeh, Wilson and Gerdes2020; Andersen et al. Reference Andersen, Elburg and Magwaza2019 a). The succession is overlain by younger sedimentary rocks of the Palaeoproterozoic Waterberg Group of the Middelburg Basin (Barker et al. Reference Barker, Brandl, Callaghan, Erikson, van der Neut, Johnson, Anhaeusser and Thomas2006) and the Phanerozoic Karoo Supergroup of the Main Karoo Basin (Johnson et al. Reference Johnson, van Vuuren, Visser, Cole, Wickens, Christie, Roberts, Brandl, Johnson, Anhaeusser and Thomas2006), which are preserved as erosional remnants in the eastern part of the area of the map in Figure 2b.

The Magaliesberg and Rayton formations were deposited some time before emplacement of the Bushveld complex, of which they form the intrusion floor (Cawthorn et al. Reference Cawthorn, Eales, Walraven, Uken, Watkeys, Johnson, Anhaeusser and Thomas2006). The age of deposition is somewhat controversial: Schröder et al. (Reference Schröder, Beukes and Armstrong2016), Beukes et al. (Reference Beukes, de Kock, Vorster, Ravhura, Frei, Gumsley and Harris2019) and Andersen et al. (Reference Andersen, Elburg and Magwaza2019 a) have suggested a maximum depositional age of c. 2.2 Ga for the Magaliesberg Formation, based on the youngest, major age fraction of detrital zircons observed, whereas Zeh et al. (Reference Zeh, Wilson and Ovtcharova2016, Reference Zeh, Wilson and Gerdes2020) interpreted zircon ages as young as 2080 Ma as protosource ages unmodified by post-depositional metamorphic resetting, and therefore prefer a correspondingly younger depositional age. No detrital zircon data have been published from the Rayton Formation.

The strata of the Transvaal Supergroup have been down-warped by emplacement of the Bushveld complex, with increasing dip towards the contact. In the south, the Vredefort meteorite impact formed a series of concentric synforms and antiforms and trust sheets around the central dome, including the Potchefstroom Syncline (Fig. 2b; Brink et al. Reference Brink, Waanders and Bisschoff1997; Therriault et al. Reference Therriault, Grieve and Reimold1997).

Emplacement of the Bushveld complex caused significant contact metamorphism in the Transvaal Supergroup, mainly in the floor rocks. Metamorphism locally reached anatectic grade in pelitic lithologies (Harris et al. Reference Harris, McMillan, Holness, Uken, Watkeys, Rodgers and Fallick2003), and sandstones are transformed to coarse-grained, glassy quartzite in the inner part of the contact aureole (Cawthorn et al. Reference Cawthorn, Eales, Walraven, Uken, Watkeys, Johnson, Anhaeusser and Thomas2006). Outside of the contact aureole (Fig. 2b), metamorphism is restricted to low-grade regional metamorphism in events dated to c. 2150 Ma and 2040 Ma (Alexandre et al. Reference Alexandre, Andreoli, Jamison and Gibson2006). Thermally induced lead loss and other effects of post-depositional metamorphism in the detrital zircon would therefore be of Palaeoproterozoic age, and no younger than the age of the Vredefort impact event at c. 2.02 Ga.

In the northeastern part of the area shown in Figure 2, the Rayton Formation is unconformably overlain by the Wilge River Formation of the Palaeoproteozoic Waterberg Group in the Middelburg Basin, comprising red-bed sandstones and conglomerates, with minor volcanic rocks (Barker et al. Reference Barker, Brandl, Callaghan, Erikson, van der Neut, Johnson, Anhaeusser and Thomas2006). These rocks post-date the Transvaal Supergroup, Rooiberg Group and Bushveld Complex, all of which have contributed material to conglomerates in the Middelburg Basin (Barker et al. Reference Barker, Brandl, Callaghan, Erikson, van der Neut, Johnson, Anhaeusser and Thomas2006).

During Carboniferous–Jurassic time, southern Africa and adjoining parts of the Gondwana supercontinent were covered by sedimentary rocks of the Karoo Supergroup, followed by lavas of the Drakensberg Group. After break-up of the supercontinent, the southeastern margin of Africa remained a topographic high. By the end of the Cretaceous Period, the inland parts of southern Africa were denuded to a high-standing, gently W-sloping erosion surface bounded by a marked escarpment to the south and east (Partridge & Maud, Reference Partridge and Maud1987; Partridge, Reference Partridge1998; Partridge et al. Reference Partridge, Botha, Haddon, Johnson, Anhaeusser and Thomas2006). In this process, the Phanerozoic cover was almost completely removed from the area of interest, so that only local relics remain; towards the east, the lower part of the Karoo succession is still intact (Fig. 2b). The erosion surface formed in this denudation cycle is partly preserved as the ‘African Surface’. Regional uplift events in the Miocene and Pliocene triggered new denudation cycles represented by two younger erosion surfaces of regional importance, of which only the older ‘Post African 1’ surface, formed in response to Miocene uplift, is of relevance for the present study (Fig. 2b). The regional denudation history implies that the units sampled for the present study have been in a surface-near position and exposed to weathering under variable climatic conditions for 20–70 Ma (Partridge, Reference Partridge1998).

3. The samples

In this study, detrital zircon has been analysed in six samples from two different exposure areas of the Pretoria Group, at different distances from the contact to the Bushveld complex (Fig. 2, Table 1). The samples were collected from roadcuts and isolated field exposures, guided by published 1:250 000 geological maps (Council for Geoscience, Pretoria). All samples were of sandstone or quartzite.

Table 1. Samples analysed for the present study

Two samples of the Rayton Formation were collected inside the limit of the contact aureole. Sample 726 is a quartzite with a distinct, pale-green colour and glassy appearance in hand specimen. Rounded detrital grains are cemented by quartz; thin rims of fine-grained chlorite around the grains are responsible for the green colour. Quartz grain boundaries are interlocking, but with sparse development of 120° triple junctions. Sample 728 is a quartz-cemented quartz-arenite. Both of the samples of the Rayton Formation contain minor (≤ 3%) chert fragments. Sample 730 of the Magaliesberg Formation was collected at the limit of the contact aureole. This sample is a partly recrystallized quartzite–quartz arenite showing graded bedding. Quartz grains are well rounded with undulose extinction and locally sutured grain contacts, with quartz cement. Samples 733, 734 and 735 come from outcrops of the Magaliesberg Formation in the central part of the Potchefstroom Syncline (Fig. 2). Sample 733 is a feldspar-free quartz arenite. Samples 734 and 735 contain up to 7% K-feldspar and minor, heavily altered lithic fragments, including chert; sample 734 also contains minor mica and apatite. All are quartz cemented, and show brown staining due to Fe-hydroxide films along grain boundaries. Zircon is a minor to accessory mineral in all of the samples (less abundant in 728 than in the other samples), occurring as well-rounded, detrital grains, without obvious post-depositional (i.e. diagenetic or metamorphic) overgrowths.

4. Methods

Samples were crushed in a steel jaw crusher, sieved to < 250 μm using a sieve with disposable cloth. Heavy mineral separates were produced by manual washing using plastic gold-washing pans. Zircon grains were picked from these separates in alcohol under a binocular microscope, mounted on two-sided adhesive tape and cast into epoxy disks, which were ground to expose the zircon grains and polished. Care was taken to produce zircon fractions that were as non-selective as possible. The zircon mounts were imaged in cathodoluminescence (CL), using a Hitachi SU5000 field emission scanning electron microscope with a Delmic Sparc-Advanced CL System at the Department of Geoscience, University of Oslo.

U–Pb and trace elements were analysed simultaneously, using a Bruker Aurora Elite quadrupole mass spectrometer with a CETAC LS 213G2+ Nd:YAG laser microprobe, also at the Department of Geosciences, University of Oslo. Trace elements were analysed using a fast scanning protocol with dwell times as given in Table 2. NIST SRM 610 was used as calibration standard, with 29Si as internal standard. Glitter software (Griffin et al. Reference Griffin, Powell, Pearson and O’Reilly2008) was used for off-line, time-resolved integration and calculation of element concentrations. Trace-element concentrations for the GJ-1 reference zircon run as an unknown are given in Table 2.

Table 2. Trace-element analyses of the GJ-1 reference zircon

U–Pb was analysed in the same ablation runs as trace elements, with real-time integration and calibration using in-house software based on Microsoft Excel. Natural zircon reference samples GJ-1 (600.5 ± 0.4 Ma, Schaltegger et al. Reference Schaltegger, Schmitt and Horstwood2015), 91500 (1065 ± 1 Ma, Wiedenbeck et al. Reference Wiedenbeck, Allé, Corfu, Griffin, Meier, Oberli, Von Quadt, Roddick and Spiegel1995), A382 (1875 ± 2 Ma, Huhma et al. Reference Huhma, Mänttäri, Peltonen, Kontinen, Halkoaho, Hanski, Hokkanen, Hölttä, Juopperi, Konnunaho, Layahe, Luukkonen, Pietikäinen, Pulkkinen, Sorjonen-Ward, Vaasjoki and Whitehouse2012) and OGC (also called OG1; 3465.4 ± 0.6 Ma, Stern et al. Reference Stern, Bodorkos, Kamo, Hickman and Corfu2009) were used for standardization, and the U/Pb ratio was internally standardized by 91Zr/29Si. 235U used for geochronology was calculated from 238U, assuming 238U/235U = 138.77 (e.g. Ludwig, Reference Ludwig2012). Raw data were reduced using an interactive spreadsheet program written in Visual Basic for Microsoft Excel. Common lead was estimated from Hg-corrected measurement of 204Pb, using a common lead composition given by the Stacey & Kramers (Reference Stacey and Kramers1975) model at the observed 206Pb/238U age of the zircon. Discordance is calculated from isotopic ratios rather than from ages (Guitreau & Blichert-Toft, Reference Guitreau and Blichert-Toft2014; Andersen et al. Reference Andersen, Elburg and Magwaza2019 a). Note that discordance values calculated by the two methods are neither identical nor linearly related.

Thirty-one analyses of an in-house reference zircon (A44, Kapinsalmi tonalite, Finland, ID-TIMS U–Pb age 2719 ± 4 Ma, Heilimo et al. Reference Heilimo, Mikkola and Halla2007) gave a weighted average 207Pb/206Pb age of 2723 ± 4 Ma (95% confidence, MSWD = 1.5).

A complete listing of U–Pb and trace-element data is given in the online Supplementary Table S1 (available at http://journals.cambridge.org/geo).

5. Results

5.a. Zircon petrography

Cathodoluminescence images of a selection of zircon grains are shown in Figures 35 to illustrate the range of variation observed in the samples studied. Detrital zircon grains are rounded, short-prismatic to elongated grains, and grain fragments. Many of the grain fragments show rounded edges cross-cutting the internal zoning pattern of the zircon, indicating that they have been abraded after breaking up (e.g. 733-67 in Fig. 4), showing that the fragmentation happened prior to final deposition. However, there are also grain fragments cut by sharp edges, which were most likely broken during crushing (e.g. grain 730-47 in Fig. 3). No systematic relationship between grain morphology or internal zoning pattern and grain size was observed, but any such relationship may have been obscured by fragmentation of the larger grains.

Fig. 3. CL photomicrographs of selected zircon grains from sample SA19-730, with chondrite-normalized REE patterns. Top to bottom: percentage of common 206Pb (bd – below detection limit), percent discordance (or conc., which indicates that the grain is concordant within error), 207Pb/206Pb age after common lead correction, if any, in Ma, and concentration of Ti, in parts per million. Length of scale bars: 50 μm. These conventions also apply to Figures 4 and 5. Chondrite values used for normalization in this and other diagrams are from Boynton (Reference Boynton and Henderson1984).

Fig. 4. CL photomicrographs of selected zircon grains from sample SA19-733, with chondrite-normalized REE patterns. See Figure 3 for abbreviations.

Fig. 5. CL photomicrographs of selected zircon grains from samples SA19-726 and SA19-728, with chondrite-normalized REE patterns. See Figure 3 for abbreviations.

The internal CL structure ranges from short-wavelength–low-amplitude oscillatory zoning (730-53 in Fig. 3; 733-67, 733-93 in Fig. 4; 726-08, 728-08 in Fig. 5), typical of unmodified igneous zircon (Corfu et al. Reference Corfu, Hanchar, Hoskin and Kinney2003), to uniformly CL-dark with only ghost-like CL-brighter inner zones (733-97 in Fig. 4; 728-15 in Fig. 5), suggesting post-crystallization modification that can be related to metamorphism in the protosource or in sedimentary precursors, or to processes after deposition. Between these extremes, there are grains showing more irregular, oscillatory zoning (730-47 in Fig. 3; 733-61 in Fig. 4; 728-25 in Fig. 5), probably still a primary feature of the zircon grain; grains with oscillatory zoning whose contrast has been enhanced, in that CL-dark zones have become darker and apparently also broader (730-47 in Fig. 3; 733-01 in Fig. 4; 728-10 and 728-31 in Fig. 5); and grains with completely irregular internal variations (730-46 in Fig. 3). CL-bright domains overprinting the oscillatory zoned zircon are relatively rare (730-52 in Fig. 3), and there is no evidence of growth of new zircon that can be attributed to diagenesis metamorphism of the host sediment. These observations apply to zircon in both Magaliesberg and Rayton formations, regardless of the degree of recrystallization of the host rock.

5.b. Trace-element distributions: Ti, Y, REE, Hf, U and Th

Chondrite-normalized REE patterns are illustrated for the grains shown as examples in Figures 35. These suggest a connection between REE distribution and CL structure, in the sense that REE patterns typical for unaltered magmatic zircon (low LREE, continuously increasing patterns towards Yb and Lu, positive Ce anomaly and negative Eu anomaly; Belousova et al. Reference Belousova, Griffin, O’Reilly and Fisher2002; Hoskin & Schaltegger, Reference Hoskin and Schaltegger2003) are restricted to grains that show reasonably well preserved oscillatory zoning patterns. The levels of light and middle REEs increase with increasing disturbance of the CL zoning pattern and, in grains with the most severely modified structure in CL images, the REE patterns are nearly flat in the middle to heavy REE range. These grains can show concentrations above 104 chondrites for the most extreme cases, and some grains even show a broad maximum in the range of Sm to Dy. These features are representative of the whole set of analysed grains, as is more conveniently illustrated in element versus element plots of the pooled dataset shown in Figure 6.

Fig. 6. Correlations with Ti of (a, b) REE Y; (c) Hf, Th and U; (d) chondrite-normalized Yb/Sm and Yb/Dy ratios (chondrite values from Boynton, Reference Boynton and Henderson1984) and (d) Lu/Hf ratio; (e) 204Pb (as proxy for common lead); and (f) U–Pb discordance.

Titanium has a range of concentrations from near the detection limit at c. 2 to 4890 ppm. Y and REE concentrations are positively correlated with Ti (Fig. 6a, b), with the largest relative variation (over 4–5 orders of magnitude) for Nd and Pr (La falls under the detection limit for some grains). Th and U are also positively correlated with Ti, whereas Hf remains close to 1.5 ×104 ppm over the whole range of Ti (Fig. 6c). Because of the increasing trend in heavy REEs, the Lu/Hf ratio also increases with increasing Ti (Fig. 6d). In contrast, the chondrite-normalized Yb/Sm and Yb/Dy ratios decrease with increasing Ti, approaching or falling below 1.0 at high Ti. This reflects the tendency towards flat patterns in the middle to heavy REEs seen in some of the examples in Figures 35; patterns with (Yb/Dy)CH < 1.0 have a maximum in the Dy-range. 204Pb ranges from below detection limit (< 0.1 ppm) to c. 10 ppm, and is positively correlated with Ti, although less so than the REE (Fig. 6e). The Th/U ratio varies from 0.16 to 4.7, with a huge majority of grains having ratios between 0.3 and 3, generally increasing with increasing concentrations (Fig. 7). This is a range in which compositions of igneous and metamorphic zircon overlap (Kirkland et al. Reference Kirkland, Smithies, Taylor, Evans and McDonald2015; Yakymchuk et al. Reference Yakymchuk, Kirkland and Clark2018).

Fig. 7. Plot of U versus Th concentration of the detrital zircon in the present study, compared with lines of constant Th/U ratio, and fields of compiled data from zircon in magmatic and metamorphic rocks (data from Kirkland et al. Reference Kirkland, Smithies, Taylor, Evans and McDonald2015 and Yakymchuk et al. Reference Yakymchuk, Kirkland and Clark2018, respectively). Filled circles: analyses with 206Pb/204Pb > 2000.

5.c. U–Pb systematics

The unfiltered, common-lead-corrected U–Pb data show a range from mildly inversely discordant to almost 100% normally discordant (Fig. 6f). The discordant grains spread widely in the concordia diagram, but with a concentration of points along a lead-loss line from c. 2200 Ma to zero (Fig. 8a). Analyses that show a 204Pb signal above the background level defined by common-lead-free reference zircons have been corrected for common lead using an average crustal lead composition according to Stacey & Kramers (Reference Stacey and Kramers1975) at the 206Pb/238U age of the zircon. This composition may not be optimal for correction for unsupported lead in zircon, causing bias towards low ages at corrections for more than 0.2–0.5% common 206Pb (Andersen et al. Reference Andersen, Elburg and Magwaza2019 a). The cumulative age distribution (Fig. 8b) of the unfiltered, common-lead-corrected dataset (297 grains) shows minimum ages of c. 2000 Ma, and a continuously increasing trend towards older, early Palaeoproterozoic and Archaean ages, with a poorly defined age fraction in the 2200–2400 Ma range. Some highly discordant grains with 206Pb/204Pb ratios below 100 give spurious ages above 4 Ga (online Supplementary Table S1); these are completely dominated by common lead that cannot be adequately corrected for, and carry no age significance. Excluding data more than 10% discordant reduces the useful dataset to 78 grains, with a major age fraction in the range 2200–2400 Ma and a smaller, late Mesoarchaean fraction at 2900–2950 Ma. A ‘tail’ of young ages between 2000 Ma and 2200 Ma persists. These apparently young grains have been corrected for common lead, and their 207Pb/206Pb ages are most likely affected by bias due to common-lead correction (cf. Andersen et al. Reference Andersen, Elburg and Magwaza2019 a). Using 206Pb/204Pb > 2000 as a limit instead of discordance (Andersen et al. Reference Andersen, Elburg and Van Niekerk2019 b) reduces the dataset further, to 47 useful analyses. The width of the main age fraction is reduced to 2225–2350 Ma, and a second Neoarchaean fraction at 2750–2800 Ma is indicated. The ‘tail’ of ages < 2200 Ma disappears.

Fig. 8. (a) Concordia diagram showing common-lead-corrected analyses of 297 detrital zircon grains from the six samples. (b) Empirical, cumulative distribution curves constructed for the full dataset, and the data after two different data filtering methods (10% discordance and 206Pb/204Pb = 2000). Shaded background represents three age fractions that can be discerned in the filtered data.

The relationship between 207Pb/206Pb age and U concentration is illustrated in Figure 9a, compared with 0, 25, 50, 75 and 100 percentiles in the granitic zircon data of Belousova et al. (Reference Belousova, Griffin, O’Reilly and Fisher2002), and to critical limits of present-day D α as a function of zircon age at 3.5 × 10–19 α/g and 8.0 × 10–19 α/g. All but six of the analysed grains plot below the 75 percentile of the concentration range of granitic zircon, but a very significant proportion will still have D α > 3.5 × 10–19 α/g , indicating that they will be significantly radiation damaged. This can be further illustrated by the cumulative distribution of D α (0) (Fig. 9b), which suggests that 68% of the zircon grains have passed the percolation limit and 39% are completely metamict. The figure also shows that the structural damage due to U and its radioactive daughters exceeds that of Th and its daughters by an order of magnitude.

Fig. 9 (a) Uranium concentration of detrital zircon plotted against common-lead-corrected 207Pb/206Pb age. Broken, horizontal lines are percentile values for the uranium concentration of igneous zircon in granitic rocks, according to Belousova et al. (Reference Belousova, Griffin, O’Reilly and Fisher2002). The 0 percentile is the minimum, 100 percentile the maximum and 50 percentile the median. Contours of D α (0) as a function of zircon age are shown for values of 3.5 × 1018 α/g and 8 × 1018 α/g, corresponding to zircon that will be at the percolation point, and to those that are fully metamict at the present time. (b) The distribution of D α (0) of the pooled set of detrital zircon in this study, calculated from Equation (1) using common-lead-corrected 207Pb/206Pb ages for t 1 and the observed U and Th concentrations. Note the concentration from U and its decay series significantly exceeds that of Th. A total of 68% of the zircon will have passed the percolation point, and 39% will be completely metamict.

6. Discussion

Detrital zircon in the samples of Magaliesberg and Rayton formations analysed in this study has suffered structural damage and disturbance to the U–Pb isotope system. Zircon of this state of preservation has limited geochronological value, and the only observations of that kind to be made with any confidence from the dataset is that the overall age distribution pattern resembles those previously reported from the Pretoria Group (Schröder et al. Reference Schröder, Beukes and Armstrong2016; Zeh et al. Reference Zeh, Wilson and Ovtcharova2016, Reference Zeh, Wilson and Gerdes2020; Andersen et al. Reference Andersen, Elburg and Magwaza2019 a; Beukes et al. Reference Beukes, de Kock, Vorster, Ravhura, Frei, Gumsley and Harris2019). Dating of deposition or provenance identification are not the concern of this study, however. The observation that grains younger than 2200 Ma are removed by the common-lead-based data filter (Section 5.c) suggests that apparent 207Pb/206Pb ages less than 2200 Ma reflect bias induced by common lead correction (Andersen et al. Reference Andersen, Elburg and Magwaza2019 a) rather than a Palaeoproterozoic metamorphic overprint.

6.a. Primary and secondary features of the detrital zircon

Elevated Ti concentration is a characteristic feature of altered zircon (Belousova et al. Reference Belousova, Griffin, O’Reilly and Fisher2002; Bell et al. Reference Bell, Boehnke, Barboni and Harrison2019). In the present samples, the Ti concentration varies over almost 5000 ppm. The primary concentration of Ti in magmatic zircon is a function of temperature and TiO2 activity (Watson et al. Reference Watson, Wark and Thomas2006), and the maximum temperatures calculated from the observed Ti concentrations would be of the order of 2000°C, which is clearly unrealistic for zircon-fertile, felsic magmatic protosource rock. This indicates that the variation in Ti is indeed due to alteration rather than to primary differentiation in the protosource(s) of detrital zircon. The lack of correlation between Ti and Hf (Fig. 6c) indicates that the Hf concentration even in heavily altered zircon has not been affected. This suggests that the Hf concentration is a feature inherited from the protosource, perhaps the only parameter analysed in this study that is preserved in the most heavily altered detrital zircon grains in these samples.

In studies of the trace-element chemistry of ancient detrital zircon in the Jack Hills conglomerate, and suites of magmatic zircon from granites of different ages, Bell et al. (Reference Bell, Boehnke and Harrison2016, Reference Bell, Boehnke, Barboni and Harrison2019) found that the sum of un-normalized ratios (Dy/Sm) + (Dy/Nd) (their LREE-I parameter) was a convenient index to distinguish primary magmatic zircon from zircon that has undergone fluid-induced alteration. Zircon with low values typically shows other evidence of secondary disturbance (high Ti, Fe, Mn, Th/U). A distinct change of slope in plots of ‘foreign’ elements versus (Dy/Sm) + (Dy/Nd) in their data at c. 50 suggested that this could be a useful limit between zircon grains that have retained their primary composition and those that have been significantly modified after crystallization. For the zircons in our study, trends in plots of, for example, U and Ti versus (Dy/Sm) + (Dy/Nd) (Fig. 10a) show a change in slope similar to that observed by Bell et al. (Reference Bell, Boehnke, Barboni and Harrison2019), but at a much lower (Dy/Sm) + (Dy/Nd) value (at c. 8, Fig. 10a), for reasons that may be related to differences in conditions (temperature, composition of fluid) or extent of alteration between the two studies. The turnover value of 8 corresponds to maximum concentrations of c. 300 ppm U and 40 ppm Ti (yielding a Ti-in-zircon temperature of c. 900°C), but with quite considerable overlap between the altered and unaltered groups, especially for U. Again, Hf concentrations show no systematic variation with (Dy/Sm) + (Dy/Nd).

Fig. 10. (a) Variation of Ti, U and Hf concentration with the sum of un-normalized ratios (Dy/Sm) + (Dy/Nd), proposed as an alteration index for zircon by Bell et al. (Reference Bell, Boehnke, Barboni and Harrison2019). See text for further explanation. (b) Variation of Th/U with (Dy/Sm) + (Dy/Nd). The shaded field is limited by Th/U ratios of 0.3 and 1.0.

The Th/U ratio shows a more complex behaviour. A large proportion of grains have Th/U between 0.3 and 1, regardless of (Dy/Sm) + (Dy/Nd) (Fig. 10b). For (Dy/Sm) + (Dy/Nd) > 8, there are only few grains with Th/U > 1, up to a maximum value of c. 2.5. In contrast, a significant number of grains with lower (Dy/Sm) + (Dy/Nd) have higher Th/U ratios, ranging to 3 and above.

Grains showing (Dy/Sm) + (Dy/Nd) > 8 have REE concentrations in the lower part of the overall range (Fig. 11), and these grains do not show any of the anomalous features with flat middle to heavy REE patterns, maxima in the middle REE range, and at least traces of positive Ce and negative Eu anomalies. The REE distributions of these grains are likely to reflect the composition of a magmatic protosource rock. In contrast, grains with (Dy/Sm) + (Dy/Nd) < 8 have high REE concentrations (up to > 4 × 104 chondrites in the heavy to middle REE range) and anomalous distribution patterns (see also Figs 35). REE patterns that are flat in the Gd to Lu range, and even relatively depleted in heavy REEs, are well known from metamorphic zircon that has grown in the presence of garnet, but such zircon will generally have heavy REE levels of 100 times chondrite or less, with distinct, positive Ce and negative Eu anomalies, and depletion in La, Pr and Nd (e.g. Hoskin & Schaltegger, Reference Hoskin and Schaltegger2003; Whitehouse & Platt, Reference Whitehouse and Platt2003; Skublov et al. Reference Skublov, Berezin and Bereshnaya2012; Johnson et al. Reference Johnson, Clark, Taylor, Santosh and Collins2015; Jiao et al. Reference Jiao, Fitzsimons and Guo2017). None of these features are seen in the present high-REE zircon. Partitioning of heavy and middle REE into xenotime during Palaeoproterozoic diagenesis or metamorphism (e.g. Rasmussen et al. Reference Rasmussen, Fletcher and Muhling2011) may cause a flattening of the heavy REE pattern in coexisting zircon, but would also be expected to cause a depletion rather than the observed increase in middle to heavy REE concentrations.

Fig. 11. Summary of chondrite-normalized REE patterns of the detrital zircon from the present study. To avoid clutter, the total variation is indicated by grey bars only (see Figs 35 for examples of actual patterns). The field of variation of detrital zircon with (Dy/Sm) + (Dy/Nd) > 8 is outlined by minimum, median and maximum lines. This group also shows a dominance of heavy over middle REEs as expected for magmatic zircon (e.g. Hoskin & Schaltegger, Reference Hoskin and Schaltegger2003). Chondrite concentrations according to Boynton (Reference Boynton and Henderson1984).

6.b. The origin of U–Pb discordance: lead loss or uranium gain

Normal discordance in zircon can in principle have four different reasons: loss of radiogenic lead, gain of uranium, incorporation of common lead and accidental mixing of concordant domains of different ages. The last of these is an analytical artefact that can normally be avoided when using laser ablation inductively coupled plasma mass spectrometry (LA-ICP-MS) in time-resolved mode, and which will not be considered further. The present data have been corrected for common lead where required, so the third mechanism can also be disregarded, although the danger of correction-induced bias towards low ages exists, as discussed in Section 5.c. The preferred interpretation of normal discordance has always been loss of radiogenic lead, which may be induced by thermal overprint or interaction with fluids (e.g. Metzger & Krogstad, Reference Metzger and Krogstad1997; Geissler et al. Reference Geissler, Schaltegger and Tomaschek2003). Low-temperature weathering processes may induce loss of radiogenic lead from radiation-damaged zircon (Stern et al. Reference Stern, Goldich and Newell1966; Pidgeon et al. Reference Pidgeon, Nemchin and Whitehouse2017). On the other hand, Pidgeon et al. (Reference Pidgeon, Nemchin and Whitehouse2017, Reference Pidgeon, Nemchin, Roberts, Whitehouse and Belluci2019) found evidence of weathering-induced U–Pb discordance in metamict zircon that must have been due to U introduction rather than to lead loss. At surface-near conditions, uranium is highly mobile as the uranyl ion (UO2 2+) (Murphy & Shock, Reference Murphy and Shock1999). Hexavalent uranium can form its own minerals (e.g. uranyl phosphates, Dal Bo et al. Reference Dal Bo, Hatert, Mees, Philippo, Baijot and Fontaine2016), and can be absorbed by metamict zircon or precipitate as secondary minerals along fractures in zircon (Pidgeon et al. Reference Pidgeon, Nemchin and Whitehouse2017, Reference Pidgeon, Nemchin, Roberts, Whitehouse and Belluci2019). If a concordant zircon with 207Pb/206Pb age t acquires D normal discordance in a recent process, the percentage discordance is given by

(3) $$D\% = 100\left( {{{{{{(^{206}}{\rm{Pb}}{/^{238}}{\rm{U}})}_{observed}}} \over {{e^{{\lambda _8}t}} - 1}} - 1} \right)$$

where λ 8 is the decay constant of 238U and ${{e^{{\lambda _8}t}} - 1}$ is the 206Pb/238U ratio of a concordant zircon with age t. Since the change of the 206Pb/238U ratio can result from reduction of radiogenic lead content by a factor x = 206Pbafter/206Pbbefore, or by an increase in the uranium concentration by a factor y = 238Uafter/238Ubefore, a discordance of D% can be expressed by either

(4) $$D\% = 100(x - 1)\,or\,D\% = 100\left( {{\raise0.5ex\hbox{$\scriptstyle 1$}\kern-0.1em/\kern-0.15em\lower0.25ex\hbox{$\scriptstyle y$}} - 1} \right)$$

depending on which process is responsible. Although increases of uranium concentration by up to an order of magnitude in the most altered zircon grains are permitted by the present data (Fig. 10a), which can generate up to 90% of normal discordance, the inverse relationship between x and y makes uranium gain less efficient than lead loss to generate such high degrees of normal discordance. Most likely, both U gain and Pb loss have contributed to the discordance pattern of the detrital zircon studied here.

6.c. Conditions of alteration

The increase in concentration of any element in altered detrital zircon requires input of the element from some source external to the zircon itself. Titanium is a high-field-strength element that has low solubility in aqueous solutions under most low-pressure and -temperature conditions (van Baalen, Reference van Baalen1993), and will tend to behave as a locally immobile element during weathering of detrital Ti-minerals such as titanite (Tilley & Eggleton, Reference Tilley and Eggleton2005). Inorganic ligands other than fluoride have negligible effect on titanium solubility at low temperature (van Baalen, Reference van Baalen1993). Elevated fluorine concentration may be important in hydrothermal systems, but is unrealistic for most for near-surface weathering scenarios. On the other hand, titanium shows enhanced low-temperature solubility at low pH, and in the presence of organic ligands (Cornu et al. Reference Cornu, Lucas, Lebon, Ambrosi, Luizão, Rouiller, Bonnay and Neal1999). The excess Ti is likely to be absorbed into the amorphous, metamict zircon, or precipitated as poorly crystalline Ti- or Fe-Ti oxides or hydroxides along fractures, as was envisaged for other non-formula elements in low-temperature altered zircon by Pidgeon et al. (Reference Pidgeon, Nemchin, Roberts, Whitehouse and Belluci2019).

Th has generally been regarded as immobile in low-temperature aqueous fluids (e.g. Braun et al. Reference Braun, Pagel, Herbillon and Rosin1992), but there is evidence for at least local remobilization of Th during weathering under hot and humid conditions (Braun et al. Reference Braun, Ngoupayou, Viers, Dupre, Bedimo, Boeglin, Robain, Nyeck, Freydier, Nkamdjou, Rouiller and Muller2005; Du et al. Reference Du, Rate and Gee2012). At pH < 7, Th forms stable, soluble complexes with phosphate-, sulphate- and organic ligands, which enhances its solubility under acid conditions (Langmuir & Herman, Reference Langmuir and Herman1980). Pidgeon et al. (Reference Pidgeon, Nemchin and Whitehouse2017, Reference Pidgeon, Nemchin, Roberts, Whitehouse and Belluci2019) found that Th can indeed be introduced into radiation-damaged zircon during weathering, and that introduction of Th and U may be at least partly independent processes. This is also indicated by our data, with a general increase in Th/U seen in the more altered zircon, suggesting that Th is introduced to a greater extent than U.

During Late Cretaceous time, the bedrock below the African Surface was deeply weathered under hot, humid and acidic (pH c. 4) conditions (Partridge & Maud, Reference Partridge and Maud1987; Partridge, Reference Partridge1998; Partridge et al. Reference Partridge, Botha, Haddon, Johnson, Anhaeusser and Thomas2006). These are conditions that would favour local mobility of Ti and Th. Weathering has continued to the present day, under variable temperature and humidity conditions. The increase in Ti seen in the altered zircon may be due to millimetre- to centimetre-scale mobilization of Ti released by leaching of detrital titanite and ilmenite during deep weathering in periods of favourable climatic conditions during Cretaceous and Cenozoic time. The accompanying enrichment in light to middle REEs, Y, U and Th is most likely due to release by leaching of detrital apatite and possibly monazite in the same process.

6.d. Implications for detrital zircon geochronology

The U–Pb and trace-element characteristics of detrital zircon in the samples of Magaliesberg and Rayton formation analysed in this study are largely consequences of Cretaceous to recent alteration induced by weathering. Less than 20% of the grains analysed have retained a reliable memory of their protosource. A data-filtering routine based on a maximum level of common lead contamination removes noise from the 207Pb/206Pb age distribution, so that distinct age fractions can be identified. Alternative trace-element-based parameters may be used to filter data, using a minimum value for (Dy/Sm) + (Dy/Nd) as suggested by Bell et al. (Reference Bell, Boehnke, Barboni and Harrison2019, but with a limit at 8 in the present data), D α (0) < 3.5 × 1018 α/g, or a maximum concentration limit for non-structural elements (e.g. Ti ≤ 40 ppm). These routines will remove grains whose trace-element chemistry has been altered, but will still allow grains with disturbed U–Pb systematics to pass, and will not reveal the pattern of age fractions suggested by the 206Pb/204Pb filtered data (Fig. 12). However, it can be a viable screening for datasets where 204Pb cannot be reported because of high mercury backgrounds during analysis.

Fig. 12. Comparison of the effect of different data filters on the detrital zircon U–Pb data. (a) Concordia diagram including ±10% discordance contours. Only points that have passed the different data filters are shown; note that points that have passed more than one filter are shown by superimposed signatures. (b) Cumulative age distribution curves for 207Pb/206Pb ages, with filters and remaining point numbers as indicated.

As was pointed out by Pidgeon et al. (Reference Pidgeon, Nemchin and Whitehouse2017), the possibility that U has been introduced by late processes has the implication that the D α (0) calculated from observed concentrations by Equation (1), and hence also the degree of metamictization estimated from Equation (2), may overestimate the real alpha dose and radiation damage of the crystal structure. Uranium added to an old zircon in Cenozoic time has only been able to induce minor radiation damage on top of what has been accumulated throughout the lifetime of the crystal at the lower, primary concentrations of the radioactive elements. If so, the calculated D α (0) value would be less useful as a filtering criterion for detrital zircon data. Furthermore, calibration methods for LA-ICP-MS that use present-day U and Th concentrations or counting rates to compensate for radiation damage of the crystal structure (e.g. Sliwinski et al. Reference Sliwinski, Guillong, Liebske, Dunkl, von Quadt and Bachmann2017) may overestimate damage and thus introduce bias. Some studies have found that the degree of radiation damage in zircon estimated from present-day U and Th concentrations exceeds the structural damage that can be measured by Raman spectroscopy, which is commonly interpreted as evidence for late annealing of the crystal structure, even where there is no other evidence of thermal overprint (e.g. Wang et al. Reference Wang, Coble, Valley, Shu, Kitajima, Spicuzza and Sun2014). This may instead be a consequence of overestimation of D α due to late introduction of U and Th. Although perhaps of less importance for detrital zircon studies, it must be noted that weathering-related change in U and Th contents also invalidates the common Pb-correction routine described by Andersen (Reference Andersen2002).

The problems outlined here affect detrital zircon in old sedimentary rocks that have been exposed to chemical weathering under climatic conditions that favour element mobility. This would apply to any Precambrian sandstone in areas undergoing tropical or sub-tropical weathering today, or which has been exposed to such weathering in the past. The present samples were taken from natural exposures and roadcuts. As long as weathering extends far below the natural surface (several tens of metres in part of South Africa; Partridge & Maud, Reference Partridge and Maud1987), surface or near-surface samples are less useful for detrital zircon geochronology than drillcore and mine samples from well below the weathered zone. Data from surface samples should therefore be interpreted with great care. Trace-element data may provide information on the extent of alteration, and should therefore be included in analytical protocols whenever possible, but may not provide a sufficiently effective data filter to remove the effects of late weathering.

Hf isotope data have not been considered in the present study. The most significant effect on time-corrected 176Hf/177Hf ratios or epsilon-Hf values is likely to be caused by shifts in the U–Pb age (Guitreau & Blichert-Toft, Reference Guitreau and Blichert-Toft2014). As a result of the elevated Hf concentration, and its apparent resistance to the late processes (Figs 6c, 10a), weathering is unlikely to affect the present-day 176Hf/177Hf ratio. However, the change in Lu/Hf ratio due to REE gain in the most severely affected zircon grains (Fig. 6d) suggests that time-corrected 176Hf/177Hf and epsilon-Hf(t) values may be modified, even when the true (i.e. protosource) age of the zircon can be estimated.

7. Conclusions

Sandstones and contact metamorphic quartzites belonging to the Magaliesberg and Rayton formations of the upper part of the Palaeoproterozoic Pretoria Group have been in a near-surface position since the Cretaceous Period, and exposed to chemical weathering under variable climatic conditions for 20–70 Ma. Whereas fully crystalline zircon has been able to survive such exposure without modification of its primary uranium–lead and trace-element characteristics, radiation-damaged zircon has been severely affected, causing increasing U–Pb discordance, increasing content of common lead and non-structural trace elements, and changing REE distribution patterns that are strongly enriched and flat to concave downwards in the middle to heavy REE range. U and Th concentrations were increased in the process, but the two elements are at least partly decoupled, resulting in a wide range of Th/U in zircon that has been affected. Late increases in U and Th have the additional consequence that the accumulated alpha dose that can be estimated from observed element concentrations is likely to overestimate the real dose absorbed through the lifetime of the zircon.

Only c. 15% of the detrital grains in the samples analysed in this study have escaped severe disturbance of their U–Pb and trace-element system. The use of data filters to such detrital zircon datasets is necessary. This study suggests that the conventional, discordance-based data filter, and alternatives based on trace-element analyses, are not sufficiently effective. Filtering the analyses on 206Pb/204Pb ratio or common lead content prior to correction works better, but leads to severe reduction of the number of useful grains in the dataset.

To achieve qualitative representativity of detrital zircon data, random sampling of zircon grains is commonly accepted as being necessary. This has the consequence that analysis of radiation-damaged zircon cannot be avoided. The results of this study raise doubts about the applicability of data from randomly sampled grains in provenance analysis of ancient sedimentary successions that have been exposed to weathering under hot and humid conditions. An alternative approach using selective sampling and/or pre-treatment of the zircon separates has the consequence that qualitative representativity is sacrificed, also causing a risk of biased interpretations.

Supplementary material

To view supplementary material for this article, please visit https://doi.org/10.1017/S001675682100114X

Acknowledgements

This study has received support from the University of Johannesburg through a DVP grant to TA; from the Department of Geosciences, University of Oslo through the departmental “Småforsk” funding program; and from Orlaug Merkesdal’s foundation. We thank Dr Magnus Kristoffersen for support in the ICPMS laboratory, and Siri Simonsen for SEM-CL imaging. MAE acknowledges the NRF for IPRR grant 119297. We thank Dr Lily Jackson for editorial handling, and reviewers Dr Udo Zimmermann and Dr Elisabeth Bell for helpful comments.

Declaration of interest

None

References

Alexandre, P, Andreoli, MAG, Jamison, A and Gibson, RL (2006) 40Ar/39Ar age constraints on low-grade metamorphism and cleavage development in the Transvaal Supergroup (central Kaapvaal craton, South Africa): implications for the tectonic setting of the Bushveld Igneous Complex. South African Journal of Geology 109, 393410.CrossRefGoogle Scholar
Andersen, T (2002) Correction of common lead in U-Pb analyses that do not report 204Pb. Chemical Geology 192, 5979.CrossRefGoogle Scholar
Andersen, T, Elburg, M and Magwaza, B (2019a) Sources of bias in detrital zircon geochronology: discordance, concealed lead loss and common lead correction. Earth-Science Reviews 197, 102899. doi:10.1016/j.earscirev.2019.102899.CrossRefGoogle Scholar
Andersen, T, Elburg, MA and Van Niekerk, HS (2019b) Detrital zircon in sandstones from the Palaeoproterozoic Waterberg and Nylstroom basins, South Africa: provenance and recycling. South African Journal of Geology 122, 7996.CrossRefGoogle Scholar
Balan, E, Neuville, DR, Trocellier, P, Fritsch, E, Muller, J-P and Calas, G (2001) Metamictization and chemical durability of detrital zircon. American Mineralogist 86, 1025–33.CrossRefGoogle Scholar
Barker, OB, Brandl, G, Callaghan, CC, Erikson, PG and van der Neut, M (2006) The Soutpansberg and Waterberg groups and the Blouberg formation. In The Geology of South Africa (eds Johnson, MR, Anhaeusser, CR and Thomas, RJ), pp. 301–18. The Geological Society of South Africa, Johannesburg/Council for Geoscience, Pretoria.Google Scholar
Bell, EA, Boehnke, P, Barboni, M and Harrison, TM (2019) Tracking chemical alteration in magmatic zircon using rare earth element abundances. Chemical Geology 510, 5671.CrossRefGoogle Scholar
Bell, EA, Boehnke, P and Harrison, TM (2016) Recovering the primary geochemistry of Jack Hills zircons through quantitative estimates of chemical alteration. Geochimica et Cosmochimica Acta 191, 187202.CrossRefGoogle Scholar
Belousova, EA, Griffin, WL, O’Reilly, SY and Fisher, NI (2002) Igneous zircon: trace element composition as an indicator of source rock type. Contributions to Mineralogy and Petrology 143, 602–22.CrossRefGoogle Scholar
Beukes, NJ, de Kock, MO, Vorster, C, Ravhura, LG, Frei, D, Gumsley, AP and Harris, C (2019) The age and country rock provenance of the Molopo Farms Complex: implications for Transvaal Supergroup correlation in southern Africa. South African Journal of Geology 122, 3956. doi: 10.25131/sajg.122.0003 CrossRefGoogle Scholar
Bindeman, IN, Schmitt, AK, Lundstrom, CC and Hervig, RL (2018) Stability of zircon and its isotopic ratios in high-temperature fluids: Long-term (4 months) isotope exchange experiment at 850°C and 50 MPa. Frontiers in Earth Science 6, Article 59, doi: 10.3389/feart.2018.00059 CrossRefGoogle Scholar
Black, LP (1987) Recent Pb loss in zircon: a natural or laboratory-induced phenomenon? Chemical Geology (Isotope Geoscience Section) 65, 2533.CrossRefGoogle Scholar
Boynton, WV (1984) Cosmochemistry of the rare earth elements; meteorite studies. In Rare Earth Element Geochemistry (ed. Henderson, P), pp. 63114. Amsterdam: Elsevier Science Publishing Co. CrossRefGoogle Scholar
Braun, J-J, Ngoupayou, JRN, Viers, J, Dupre, B, Bedimo, J-PB, Boeglin, J-L, Robain, H, Nyeck, B, Freydier, R, Nkamdjou, LS, Rouiller, J and Muller, J-P (2005) Present weathering rates in a humid tropical watershed: Nsimi, South Cameroon. Geochimica et Cosmochimica Acta 69, 357–87.CrossRefGoogle Scholar
Braun, J-J, Pagel, M, Herbillon, A and Rosin, C (1992) Mobilization and redistribution of REEs and thorium in a syenitic lateritic profile: a mass balance study. Geochimica et Cosmochimica Acta 57, 4419–34.CrossRefGoogle Scholar
Brink, MC, Waanders, FB and Bisschoff, AA (1997) Vredefort: a model for the anatomy of an astrobleme Tectonophysics 270, 83114.CrossRefGoogle Scholar
Brink, MC, Waanders, FB, Bisschoff, AA and Gay, NC (2000) The Foch Trust – Potscefstroom Fault structural stystem, Vredefort, South Africa: a model for impact related tectonic movement over a pre-existing barrier. Journal of African Earth Sciences 30, 99117.CrossRefGoogle Scholar
Cawthorn, RG, Eales, HV, Walraven, F, Uken, R and Watkeys, MK (2006) The Bushveld Complex. In The Geology of South Africa (eds Johnson, MR, Anhaeusser, CR and Thomas, RJ), pp. 261281. The Geological Society of South Africa, Johannesburg/Council for Geoscience, Pretoria.Google Scholar
Corfu, F, Hanchar, JM, Hoskin, PWO and Kinney, P (2003) Atlas of zircon textures. Reviews in Mineralogy and Geochemistry 53, 469500.CrossRefGoogle Scholar
Cornu, S, Lucas, Y, Lebon, E, Ambrosi, JP, Luizão, F, Rouiller, J, Bonnay, M and Neal, C (1999) Evidence of titanium mobility in soil profiles, Manaus, central Amazonia. Geoderma 91, 281–95.CrossRefGoogle Scholar
Dal Bo, F, Hatert, F, Mees, F, Philippo, S, Baijot, M and Fontaine, F (2016) Crystal structure of bassetite and saléeite: new insight into autunite-group minerals. European Journal of Mineralogy 28, 663–75.CrossRefGoogle Scholar
Dorland, HC (2004) Provenance Ages and Timing of Sedimentation of Selected Neoarchaean and Palaeoproterozoic Successions on the Kaapvaal Craton. Ph.D. thesis, Rand Afrikaans University, Johannesburg. Published thesis.Google Scholar
Du, X, Rate, AW and Gee, MAM (2012) Redistribution and mobilization of titanium, zirconium and thorium in an intensely weathered lateritic profile in Western Australia. Chemical Geology 330–332, 101–15.CrossRefGoogle Scholar
Eriksson, PG, Altermann, W, Catuneanu, O, van der Merwe, R and Bumby, AJ (2001) Major influences on the evolution of the 2.67–2.1 Ga Transvaal basin, Kaapvaal craton. Sedimentary Geology 141–142, 205–31.CrossRefGoogle Scholar
Eriksson, PG, Altermann, W and Hartzer, FJ (2006) The Transvaal Supergroup and its precursors. In The Geology of South Africa (eds Johnson, MR, Anhaeusser, CR and Thomas, RJ), pp. 237–60. The Geological Society of South Africa, Johannesburg/Council for Geoscience, Pretoria.Google Scholar
Ewing, RC, Meldrum, A, Wang, L, Weber, WJ and Corrales, LR (2003) Radiation effects in zircon. Reviews in Mineralogy and Geochemistry 53, 387435.CrossRefGoogle Scholar
Geissler, T, Schaltegger, U and Tomaschek, F (2003) Re-equilibration of zircon in aqueous fluids and melts. Elements 3, 4350.CrossRefGoogle Scholar
Griffin, WL, Belousova, EA, Shee, SR, Pearson, NJ and O’Reilly, SY (2004) Archean crustal evolution in the northern Yilgarn Craton: U–Pb and Hf-isotope evidence from detrital zircons. Precambrian Research 131, 231–82.CrossRefGoogle Scholar
Griffin, WL, Powell, WJ, Pearson, N and O’Reilly, SY (2008) GLITTER: Data reduction software for laser ablation ICP-MS. Mineralogical Association of Canada Short Course Series 40, 204–7.Google Scholar
Guitreau, M and Blichert-Toft, J (2014) Implications of discordant U-Pb ages on Hf isotope studies. Chemical Geology 385, 1725.CrossRefGoogle Scholar
Gumsley, AP, Chamberlain, KR, Bleeker, W, Söderlund, U, de Kock, MO, Larsson, ER and Bekker, A (2017) Timing and tempo of the Great Oxidation Event. Proceedings of the National Academy of Sciences of the USA 114, 1811–6. doi:10.1073/pnas.1608824114 CrossRefGoogle ScholarPubMed
Hannah, JL, Bekker, A, Stein, HJ, Markey, RJ and Holland, HD (2004) Primitive Os and 2316 Ma age for marine shale: implications for Paleoproterozoic glacial events and the rise of atmospheric oxygen. Earth Planet Science Letters 225, 4352.CrossRefGoogle Scholar
Harris, N, McMillan, A, Holness, M, Uken, R, Watkeys, M, Rodgers, N and Fallick, A (2003) Melt generation and fluid flow in the thermal aureole of the Bushveld Complex. Journal of Petrology 44, 1031–54.CrossRefGoogle Scholar
Hay, DC and Dempster, TJ (2009) Zircon alteration, formation and preservation in sandstones. Sedimentology 56, 2175–91.CrossRefGoogle Scholar
Heilimo, E, Mikkola, PO and Halla, J (2007) Age and petrology of the Kaapinsalmi sanukitoid intrusion in Suomussalmi, Eastern Finland. Bulletin of the Geological Society of Finland 79, 117–25.CrossRefGoogle Scholar
Holland, H and Gottfried, D (1955) The effect of nuclear radiation on the structure of zircon. Acta Crystallographica 9, 291300.CrossRefGoogle Scholar
Hoskin, PWO and Schaltegger, U (2003) The composition of zircon and igneous and metamorphic petrogenesis. Reviews in Mineralogy and Geochemistry 53, 2762.CrossRefGoogle Scholar
Huhma, H, Mänttäri, I, Peltonen, P, Kontinen, A, Halkoaho, T, Hanski, E, Hokkanen, T, Hölttä, P, Juopperi, H, Konnunaho, J, Layahe, Y, Luukkonen, E, Pietikäinen, K, Pulkkinen, A, Sorjonen-Ward, P, Vaasjoki, M and Whitehouse, M (2012) The age of the Archaean greenstone belts in Finland. Espoo: Geological Survey of Finland, Special Paper, 54, 74–175.Google Scholar
Jiao, S, Fitzsimons, ICW and Guo, J (2017) Paleoproterozoic UHT metamorphism in the Daqingshan Terrane, North China Craton: new constraints from phase equilibria modeling and SIMS U–Pb zircon dating. Precambrian Research 303, 208–27.CrossRefGoogle Scholar
Johnson, MR, van Vuuren, CJ, Visser, JNJ, Cole, DI, Wickens, H de V, Christie, ADM, Roberts, DL and Brandl, G (2006) Sedimentary Rocks of the Karoo Supergroup. In The Geology of South Africa (eds Johnson, MR, Anhaeusser, CR and Thomas, RJ), pp. 461–99. The Geological Society of South Africa, Johannesburg/Council for Geoscience, Pretoria.Google Scholar
Johnson, TE, Clark, C, Taylor, RJM, Santosh, M and Collins, AS (2015) Prograde and retrograde growth of monazite in migmatites: an example from the Nagercoil Block, southern India. Geoscience Frontiers 6, 373–87.CrossRefGoogle Scholar
Kamo, SL, Reimold, WU, Krogh, TE and Colliston, WP (1996) A 2.023 Ga age for the Vredefort impact event and a first report of shock metamorphosed zircons in pseudotachylite breccias and granophyre. Earth and Planetary Science Letters 144, 369–88.CrossRefGoogle Scholar
Kirkland, CL, Smithies, RH, Taylor, RJM, Evans, N and McDonald, B (2015) Zircon Th/U ratios in magmatic environs. Lithos 212–215, 397414.CrossRefGoogle Scholar
Krogh, TE (1982) Improved accuracy of U-Pb zircon ages by the creation of more concordant systems using an air abrasion technique. Geochimica et Cosmochimica Acta 46, 637–49.CrossRefGoogle Scholar
Langmuir, D and Herman, JS (1980) The mobility of thorium in natural waters at low temperture. Geochimica et Cosmochimica Acta 44, 1753–66.CrossRefGoogle Scholar
Ludwig, KR (2012) User’s Manual for Isoplot/Excel 3.75: A Geochronological Toolkit for Microsoft Excel. Berkeley: Berkeley Geochronology Center, Special Publication no. 5, 75 pp.Google Scholar
Mapeo, RBM, Armstrong, RA, Kampunzu, AB, Modisi, MP, Ramokate, LV and Modie, BNJ (2006) A ca. 200Ma hiatus between the Lower and Upper Transvaal Groups of southern Africa: SHRIMP U–Pb detrital zircon evidence from the Segwagwa Group, Botswana: Implications for Palaeoproterozoic glaciations. Earth and Planetary Science Letters 244, 113–32.CrossRefGoogle Scholar
Markwitz, V and Kirkland, CL (2018) Source to sink zircon grain shape: constraints on selective preservation and significance for Western Australian Proterozoic basin provenance. Geoscience Frontiers 9, 415–30.CrossRefGoogle Scholar
Mattinson, JM (2005) Zircon U–Pb chemical abrasion (“CA-TIMS”) method: combined annealing and multi-step partial dissolution analysis for improved precision and accuracy of zircon ages. Chemical Geology 220, 4756.CrossRefGoogle Scholar
Metzger, K and Krogstad, EJ (1997) Interpretation of discordant U-Pb zircon ages: An evaluation. Journal of Metamorphic Geology 15, 127–40.CrossRefGoogle Scholar
Moore, JM, Polteau, S, Armstrong, RA, Corfu, F and Tsikos, H (2012) The age and correlation of the Postmasburg Group, southern Africa: constraints from detrital zircon grains. Journal of African Earth Sciences 64, 919.CrossRefGoogle Scholar
Murphy, WM and Shock, EL (1999) Environmental aqueous geochemistry of actinides. Reviews in Mineralogy and Geochemistry 38, 221–53.CrossRefGoogle Scholar
Nasdala, L, Reiners, PW, Garver, JI, Kennedy, AK, Stern, RA, Balan, E and Wirth, R (2004) Incomplete retention of radiation damage in zircon from Sri Lanka. American Mineralogist 89, 219–31.CrossRefGoogle Scholar
Nasdala, L, Wenzel, M, Vavra, G, Irmer, G, Wenzel, T and Kober, B (2001) Metamictizatoin of natural zircon: accumulation versus thermal annealing of radioactivity-induced damage. Contributions to Mineralogy and Petrology 141, 125144.CrossRefGoogle Scholar
Partridge, TC (1998) Of diamonds, dinosaurs and diastrophism: 150 million years of landscape evolution in southern Africa. South African Journal of Geology 101, 167–84.Google Scholar
Partridge, TC, Botha, GA and Haddon, JG (2006) Cenozoic deposits of the interior. In The Geology of South Africa (eds Johnson, MR, Anhaeusser, CR and Thomas, RJ), pp. 585604. Geological Society of South Africa, Johannesburg/Council for Geoscience, Pretoria.Google Scholar
Partridge, TC and Maud, RR (1987) Geomorphic evolution of southern Africa since the Mesozoic. South African Journal of Geology 90, 179208.Google Scholar
Piazolo, S, Belousova, E, La Fontaine, A, Corcoran, C and Cairney, JM (2017) Trace element homogeneity from micron- to atomic scale: implication for the suitability of the zircon GJ-1 as a trace element reference material. Chemical Geology 456, 10–8.CrossRefGoogle Scholar
Pidgeon, RT, Nemchin, AA and Cliff, J (2013) Interaction of weathering solutions with oxygen and U–Pb isotopic systems of radiation-damaged zircon from an Archean granite, Darling Range Batholith, Western Australia. Contributions to Mineralogy and Petrology 166, 511–23.CrossRefGoogle Scholar
Pidgeon, RT, Nemchin, AA, Roberts, MP, Whitehouse, MJ and Belluci, JJ (2019) The accumulation of non-formula elements in zircons during weathering: ancient zircons from the Jack Hills, Western Australia. Chemical Geology 530, 119310.CrossRefGoogle Scholar
Pidgeon, RT, Nemchin, AA and Whitehouse, MJ (2017) The effect of weathering on U–Th–Pb and oxygen isotope systems of ancient zircons from the Jack Hills, Western Australia. Geochimica et Cosmochimica Acta 197, 142–66.CrossRefGoogle Scholar
Rasmussen, B, Bekker, A and Fletcher, IR (2013) Correlation of Paleoproterozoic glaciations based on U–Pb zircon ages for tuff beds in the Transvaal and Huronian Supergroups. Earth and Planetary Science Letters 382, 173–80.CrossRefGoogle Scholar
Rasmussen, B, Fletcher, IR and Muhling, JR (2011) Response of xenotime to prograde metamorphism. Contributions to Mineralogy and Petrology 162, 1259–77.CrossRefGoogle Scholar
Salje, EKH (2006) Elastic softening of zircon by radiation damage. Applied Physics Letters 89, 131902. CrossRefGoogle Scholar
Salje, EKH, Chronsch, J and Ewing, RC (1999) Is “metamictization” of zircon a phase transition? American Mineralogist 84, 1107–16.CrossRefGoogle Scholar
Schaltegger, U, Schmitt, AK and Horstwood, MSA (2015) U–Th–Pb zircon geochronology by ID-TIMS, SIMS, and laser ablation ICP-MS: recipes, interpretations, and opportunities. Chemical Geology 402, 89110.CrossRefGoogle Scholar
Schreiber, UM and Eriksson, PG (1992) The sedimentology of the post-Magaliesberg formation of the Pretoria Group, Transvaal Sequence, in the eastern Transvaal. South African Journal of Geology 95, 116.Google Scholar
Schröder, S, Beukes, NJ and Armstrong, RA (2016) Detrital zircon constraints on the tectonostratigraphy of the Paleoproterozoic Pretoria Group, South Africa. Precambrian Research 278, 362–93.CrossRefGoogle Scholar
Skublov, SG, Berezin, AV and Bereshnaya, NG (2012) General relations in the trace element composition of zircons from eclogites with implications for the age of eclogites in the Belomorian Mobile Belt. Petrology 20, 427–49.CrossRefGoogle Scholar
Sliwinski, JT, Guillong, M, Liebske, C, Dunkl, I, von Quadt, A and Bachmann, O (2017) Improved accuracy of LA-ICP-MS U-Pb ages of Cenozoic zircons by alpha dose correction. Chemical Geology 472, 821.CrossRefGoogle Scholar
Stacey, JS and Kramers, JD (1975) Approximation of terrestrial lead isotope evolution by a two-stage model. Earth Planetary Science Letters 26, 207–21.CrossRefGoogle Scholar
Stern, RA, Bodorkos, S, Kamo, S, Hickman, AH and Corfu, F (2009) Measurement of SIMS instrumental mass fractionation of Pb isotopes during zircon dating. Geostandards and Geoanalytical Research 33, 145–68.CrossRefGoogle Scholar
Stern, TW, Goldich, SS and Newell, MF (1966) Effects of weathering on the U-Pb ages of zircon from the Morton Gneiss, Minnesota. Earth and Planetary Science Letters 1, 369–71.CrossRefGoogle Scholar
Therriault, AM, Grieve, RA and Reimold, U (1997) Original size of the Vredefort structure: implications for the geological evolution of the Witwatersrand Basin. Meteoritics and Planetary Science 32, 71–7.CrossRefGoogle Scholar
Tilley, DB and Eggleton, RA (2005) Titanite low-temperature alteration and Ti mobility. Clays and Clay Minerals 53, 100–7.CrossRefGoogle Scholar
van Baalen, MR (1993) Titanium mobility in metamorphic systems: a review. Chemical Geology 110, 233–49.CrossRefGoogle Scholar
Veevers, JJ and Saeed, A (2007) Central Antarctic provenance of Permian sandstones in Dronning Maud Land and the Karoo Basin: integration of U-Pb age and TDM ages and host-rock affinity from detrital zircons. Sedimentary Geology 202, 653–76.CrossRefGoogle Scholar
Wabo, H, Humbert, F, de Kock, MO, Belyanin, G, Söderlund, U, Maré, LP and Beukes, NJ (2019) Constraining the chronology of the Mashishing Dykes from the Eastern Kaapvaal Craton in South Africa. In Dyke Swarms of the World: A Modern Perspective (eds Srivastava, RK, Ernst, RE and Peng, P), pp. 215–61. Springer Geology, doi:10.1007/978-981-13-1666-1_6.CrossRefGoogle Scholar
Walraven, F (1997) Geochronology of the Rooiberg Group, Transvaal Supergroup, South Africa. Information Circular No. 316, Economic Geology Research Unit, University of the Witwatersrand, Johannesburg, South Africa, 21 pp.Google Scholar
Wang, X-L, Coble, MA, Valley, JW, Shu, X-J, Kitajima, K, Spicuzza, MJ and Sun, T (2014) Influence of radiation damage on Late Jurassic zircon from southern China: evidence from in situ measurements of oxygen isotopes, laser Raman, U–Pb ages, and trace elements. Chemical Geology 389, 122–36.CrossRefGoogle Scholar
Watson, EB, Wark, DA and Thomas, JB (2006) Crystallization thermometers for zircon and rutile. Contributions to Mineralogy and Petrology 151, 413–33. doi: 10.1007/s00410-006-0068-5 CrossRefGoogle Scholar
Whitehouse, MJ and Platt, JP (2003) Dating high-grade metamorphism—constraints from rare-earth elements in zircon and garnet. Contributions to Mineralogy and Petrology 145, 6174.CrossRefGoogle Scholar
Wiedenbeck, M, Allé, P, Corfu, F, Griffin, WL, Meier, M, Oberli, F, Von Quadt, A, Roddick, JC and Spiegel, W (1995) Three natural zircon standards for U-Th-Pb, Lu-Hf, trace element and REE analysis. Geostandards Newsletter 19, 123.CrossRefGoogle Scholar
Williams, IS (2001) Response of detrital zircon and monazite, and their U–Pb isotopic systems, to regional metamorphism and host-rock partial melting, Cooma Complex, southeastern Australia. Australian Journal of Earth Sciences 48, 557–80.CrossRefGoogle Scholar
Willner, AP, Sindern, S, Metzger, K, Ermolaeva, T, Kramm, U, Puchkov, V and Kronz, A (2003) Typology and single grain U/Pb ages of detrital zircons from Proterozoic sandstones in the SW Urals (Russia): early time marks at the eastern margin of Baltica. Precambrian Research 124, 129.CrossRefGoogle Scholar
Yakymchuk, C, Kirkland, CL and Clark, C (2018) Th/U ratios in metamorphic zircon. Journal of Metamorphic Geology 36, 715–37.CrossRefGoogle Scholar
Zeh, A, Ovtcharova, M, Wilson, AH and Schaltegger, U (2015) The Bushveld complex was emplaced and cooled in less than one million years – results of zirconology, and geotectonic implications. Earth and Planetary Science Letters 418, 103–14.CrossRefGoogle Scholar
Zeh, A, Wilson, AH and Gerdes, A (2020) Zircon U-Pb-Hf isotope systematics of Transvaal Supergroup – Constraints for the geodynamic evolution of the Kaapvaal Craton and its hinterland between 2.65 and 2.06 Ga. Precambrian Research 345, 105760. doi:10.1016/j.precamres.2020.105760 CrossRefGoogle Scholar
Zeh, A, Wilson, AH and Ovtcharova, M (2016) Source and age of upper Transvaal Supergroup, South Africa: age-Hf isotope record of zircons in Magaliesberg quartzite and Dullstroom lava, and implications for Paleoproterozoic (2.5–2.0 Ga) continent reconstruction. Precambrian Research 278, 121.CrossRefGoogle Scholar
Zhang, M and Salje, EKH (2001) Infrared spectroscopic analysis of zircon: radiation damage and the metamict state. Journal of Physics: Condensed Matter 13, 3057–71.Google Scholar
Zhao, L, Guo, F, Fan, W, Zhang, Q, Wu, Y, Li, J and Yan, W (2016) Early Cretaceous potassic volcanic rocks in the Jiangnan Orogenic Belt, East China: crustal melting in response to subduction of the Pacific–Izanagi ridge? Chemical Geology 437, 3043.CrossRefGoogle Scholar
Zimmermann, U (2018) The provenance of selected Neoproterozoic to lower Paleozoic basin successions of Southwest Gondwana: a review and proposal for further research. In Geology of Southwest Gondwana (eds Siegesmund, S, Basei, MAS, Oyhantcabal, P and Oriolo, S), pp. 561–91. Springer International Publishing AG, Regional Geology Reviews.CrossRefGoogle Scholar
Figure 0

Fig. 1. Distribution of accumulated alpha dose at t = 0 Ma experienced by 2.2 Ga zircon with U and Th concentration distributions similar to those reported for zircon in granitic rocks by Belousova et al. (2002), represented by percentile points and a log-normal distribution compatible with these, and for a wider compositional range of igneous rocks by Kirkland et al. (2015); log-normal distribution based on data from their supplementary table A1. Dα(0) values at which zircon would have reached the percolation point (3.5 × 1018 α/g) and the complete metamictization limit (8.0 × 1018 α/g) are as given by Salje et al. (1999).

Figure 1

Fig. 2. (a) Generalized stratigraphic column of the Pretoria Group in the south-central part of the Transvaal Basin, South Africa, simplified from Eriksson et al. (2006, fig. 9). The Rooiberg Group lavas have been dated to 2061 ± 2 Ma by a lead evaporation age on zircon (Walraven, 1997), and the intrusive rocks of the Bushveld complex (BVC) by ID-TIMS U–Pb on zircon to 2056 ± 0.3 Ma by Zeh et al. (2015). Further geochronological evidence limiting time of deposition of Pretoria Group strata are: (1) Timeball Hill Formation, syn-sedimentary ashlayers, Rasmussen et al. (2013); (2) younger ashlayers in the Timeball Hill Formation also provide the currently most robust available maximum limit for the age of the Hekpoort lavas; and (3) minimum age of the Daspoort and lower part of the Silverton formations sandstone is given by an Ar–Ar age on a cross-cutting mafic–ultramafic dyke swarm (Wabo et al.2019). Sample numbers shown in parentheses (Magaliesberg and Rayton formations) refer to localities shown by circles in (b). (b) Simplified geological map of the south-central part of the Transvaal Basin, after Council of Geoscience 1: 250 000 geological mapsheets Rustenburg, Pretoria, West Rand and East Rand. The 500°C isograd of the Bushveld contact aureole (i) and the outer limit of the aureole (ii) are from Cawthorn et al. (2006). Extents of preserved African and Post African 1 surfaces are from Partridge (1998). V – Vredefort Dome, centre of the 2.02 Ga Vredefort meteorite impact; PS – axis of the Potchefstroom Syncline, from Brink et al. (2000).

Figure 2

Table 1. Samples analysed for the present study

Figure 3

Table 2. Trace-element analyses of the GJ-1 reference zircon

Figure 4

Fig. 3. CL photomicrographs of selected zircon grains from sample SA19-730, with chondrite-normalized REE patterns. Top to bottom: percentage of common 206Pb (bd – below detection limit), percent discordance (or conc., which indicates that the grain is concordant within error), 207Pb/206Pb age after common lead correction, if any, in Ma, and concentration of Ti, in parts per million. Length of scale bars: 50 μm. These conventions also apply to Figures 4 and 5. Chondrite values used for normalization in this and other diagrams are from Boynton (1984).

Figure 5

Fig. 4. CL photomicrographs of selected zircon grains from sample SA19-733, with chondrite-normalized REE patterns. See Figure 3 for abbreviations.

Figure 6

Fig. 5. CL photomicrographs of selected zircon grains from samples SA19-726 and SA19-728, with chondrite-normalized REE patterns. See Figure 3 for abbreviations.

Figure 7

Fig. 6. Correlations with Ti of (a, b) REE Y; (c) Hf, Th and U; (d) chondrite-normalized Yb/Sm and Yb/Dy ratios (chondrite values from Boynton, 1984) and (d) Lu/Hf ratio; (e) 204Pb (as proxy for common lead); and (f) U–Pb discordance.

Figure 8

Fig. 7. Plot of U versus Th concentration of the detrital zircon in the present study, compared with lines of constant Th/U ratio, and fields of compiled data from zircon in magmatic and metamorphic rocks (data from Kirkland et al.2015 and Yakymchuk et al.2018, respectively). Filled circles: analyses with 206Pb/204Pb > 2000.

Figure 9

Fig. 8. (a) Concordia diagram showing common-lead-corrected analyses of 297 detrital zircon grains from the six samples. (b) Empirical, cumulative distribution curves constructed for the full dataset, and the data after two different data filtering methods (10% discordance and 206Pb/204Pb = 2000). Shaded background represents three age fractions that can be discerned in the filtered data.

Figure 10

Fig. 9 (a) Uranium concentration of detrital zircon plotted against common-lead-corrected 207Pb/206Pb age. Broken, horizontal lines are percentile values for the uranium concentration of igneous zircon in granitic rocks, according to Belousova et al. (2002). The 0 percentile is the minimum, 100 percentile the maximum and 50 percentile the median. Contours of Dα(0) as a function of zircon age are shown for values of 3.5 × 1018 α/g and 8 × 1018 α/g, corresponding to zircon that will be at the percolation point, and to those that are fully metamict at the present time. (b) The distribution of Dα(0) of the pooled set of detrital zircon in this study, calculated from Equation (1) using common-lead-corrected 207Pb/206Pb ages for t1 and the observed U and Th concentrations. Note the concentration from U and its decay series significantly exceeds that of Th. A total of 68% of the zircon will have passed the percolation point, and 39% will be completely metamict.

Figure 11

Fig. 10. (a) Variation of Ti, U and Hf concentration with the sum of un-normalized ratios (Dy/Sm) + (Dy/Nd), proposed as an alteration index for zircon by Bell et al. (2019). See text for further explanation. (b) Variation of Th/U with (Dy/Sm) + (Dy/Nd). The shaded field is limited by Th/U ratios of 0.3 and 1.0.

Figure 12

Fig. 11. Summary of chondrite-normalized REE patterns of the detrital zircon from the present study. To avoid clutter, the total variation is indicated by grey bars only (see Figs 3–5 for examples of actual patterns). The field of variation of detrital zircon with (Dy/Sm) + (Dy/Nd) > 8 is outlined by minimum, median and maximum lines. This group also shows a dominance of heavy over middle REEs as expected for magmatic zircon (e.g. Hoskin & Schaltegger, 2003). Chondrite concentrations according to Boynton (1984).

Figure 13

Fig. 12. Comparison of the effect of different data filters on the detrital zircon U–Pb data. (a) Concordia diagram including ±10% discordance contours. Only points that have passed the different data filters are shown; note that points that have passed more than one filter are shown by superimposed signatures. (b) Cumulative age distribution curves for 207Pb/206Pb ages, with filters and remaining point numbers as indicated.

Supplementary material: File

Andersen and Elburg supplementary material

Andersen and Elburg supplementary material

Download Andersen and Elburg supplementary material(File)
File 378.9 KB